Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Rotational excitation of C2H anion in collision with H2

Insaf Toumia, Ounaies Yazidi*ab and Faouzi Najarac
aLaboratoire de Spectroscopie Atomique, Moléculaire et Application, Faculté des Sciences, Université Tunis El Manar, Tunis 2092, Tunisia
bInstitut Préparatoire aux Etudes d'Ingénieurs de Tunis El Manar, Université de Tunis El Manar, Tunis 2092, Tunisia
cInstitut Préparatoire aux Etudes d'Ingénieurs de Tunis, Université de Tunis, Tunis 1007, Tunisia

Received 21st January 2021 , Accepted 5th April 2021

First published on 13th April 2021


Abstract

The discovery of anions in the interstellar medium has shown that they are very reactive species. This gave them great importance in the modeling of the chemical and astrophysical evolution of the interstellar medium. The detection of the first anion C6H followed by the other anions C4H, C8H and CN in the interstellar medium has encouraged research on other detectable anions. The C2H anion was observed for the first time in the circumstellar envelope of IRC+10216 and in TMC-1. In these cold and low-density regions, precise modeling of the chemical and physical conditions of the observed emission lines requires knowledge of the radiative and collisional excitation rates. We present here the first new two-dimensional Potential Energy Surface (PES) for C2H–H2 interaction. Rotational excitation of the anion by collision with para-H2(jH2 = 0) is investigated. The PES is obtained in the super-molecular approach based on a single and double excitation coupled cluster method with perturbative contributions from triple excitations (CCSD(T)). In all our calculations, all atoms were described using the augmented correlation-consistent triple zeta (aug-cc-pVTZ) basis sets and bond f unctions. Fully-quantum close-coupling calculations of inelastic integral cross sections are done on a grid of collision energies large enough to ensure converged state-to-state rate coefficients for the 16 first rotational levels of C2H and for temperatures ranging from 5 to 120 K. For this collisional system, rate coefficients exhibit a strong propensity in favor of even Δj transitions.


1 Introduction

In the last two decades, anions have been intensively studied. These studies were framed by observations in the laboratory1–3 and some astrophysical observations.2,4–6 Among these species, mention is made of carbon chain anions C2nH.6 The recent discovery of carbon chain anions C2nH in interstellar and circumstellar media has been investigated by many theoretical and experimental works on these species.7,8 Their structures as well as the importance of their role in the interstellar chemistry and in gas phase ion–molecule reactions are the object of many recent studies.9–11 Although the existence of anions in astrophysical sources was first predicted theoretically and early considered in chemical models,12,13 the first negative hydrocarbon C6H was detected in 2006 (ref. 2) solving the problem of the unidentified lines discovered by Kawaguchi et al.14 The C6H identification was followed by the detection of other negatively charged species like C4H, C8H, C3N, C5N and CN.15–23 Many of these species were first detected in IRC+10216.2,15,18 Hydrocarbon anions were also discovered later in other molecular clouds.17 In this set C2H present a capital importance.

Several observations in the laboratories have justified the detection of the C2H anion. This experimental studies24 interest of the C2H anion comes from the fact that the C2H anion is the shortest in the sequence of carbon chain anions with a closed-shell electronic ground state, and the fact that gives the rest frequencies required for a radio astronomical search for this polar, astronomically plausible molecule.

From astrophysical observations, in 2007 J. Cernicharo et al.15,25 detected the 1–0 transition of the C2H anion and calculated the abundance ratio C2H/C2H = 12.5. Cernicharo et al. suggest that the 3–2 transition is strong enough to be detected. Later in 2010, the observations of M. Agùndez et al.26 found difficulties in detecting the 1–0 and 2–0 transitions of the anion. These difficulties are probably due to the very reactive natures of the C2H anion.

We report in this work a first collisional study of C2H with para-H2(jH2 = 0) at low temperatures trying to understand the particular behavior of negatively charged species during collisions and how they compare with neutral forms27,47 and the C2H anion in collision with He.48 In the next section, we will present details of the ab initio calculations of the C2H–para-H2 potential Energy Surface (PES).In Section 3, some theoretical aspects of the scattering calculations are given, state-to-state collisional cross sections will be reported, the rate coefficients and the propensity rules will be showcased. In Section 4, comparative study between the rate coefficients of C2Hpara-H2 and C2H–He will be displayed. Concluding remarks are in Section 5.

2 Potential Energy Surface

This work primarily focuses on the rotational excitation of C2H with para-H2 at low temperature; both monomers were considered as rigid species to deal with a reduced number of degrees of freedom. Accordingly, in our calculations, is considered to be linear, and both bond lengths are set at the CCSD/AV5Z theoretical values of rC–C = 1.250 Å and rC–H = 1.070 Å.28 For H2, we used the experimental bond length rH–H = 1.44876 bohr corresponding to the averaged value over the ground-state vibrational wave function for H2. The lowest bending mode of C2H ν1 equal to 502.0 cm−1,28 is well above the energy of the highest C2H rotational level considered in this work (j = 16, see Table 1). In addition, from QCT studies of rotational excitation of H2O by H2, it was shown29 that the coupling between the rotational excitation and the bending may be neglected for temperatures up to 5000 K. Moreover, it was also shown by Faure et al.,30 in the particular case of the H2O–H2 system, that the 5D-PES calculated at the experimental ground vibrational state geometry and the full 9D-PES averaged over the ground vibrational state are very similar. These results suggest that the use of a rigid body 2D-PES is sufficient as long as bending and vibrational excitations are not taken into consideration.
Table 1 Rotational Energy for the first 28 Levels of C2H system
j εj (cm−1) j εj (cm−1)
0 0.0000000    
1 2.7778579 15 333.1581888
2 8.3334960 16 377.5511276
3 16.6667590 17 424.7113668
4 27.7774142 18 474.6375867
5 41.6651509 19 527.3283902
6 58.3295811 20 582.7823022
7 77.7702390 21 640.9977702
8 99.9865812 22 701.9731640
9 124.9779866 23 765.7067757
10 152.7437566 24 832.1968200
11 183.2831149 25 901.4414338
12 216.5952076 26 973.4386763
13 252.6791032 27 1048.1865292
14 291.5337926 28 1125.6828966


Then, we define the body-fixed coordinate system in Fig. 1. The geometry of C2H–H2 collisional system with para-H2 and C2H treated as rigid rotors is then characterized by three angles (θ, θ′, ϕ) and the distance R between the center of masses of H2 and C2H (see Fig. 1). The polar angles of the C2H and para-H2 molecules with respect to R are denoted θ and θ′ respectively, while ϕ denotes the dihedral angle, which is the relative polar angle between the C2H and para-H2 bonds. For the solution of the close-coupling scattering equations, it is most convenient to expand, at each value of R, the interaction potential V(R,θ,θ′,ϕ) in angular functions. For the scattering of two linear rigid rotors, we used.31

 
image file: d1ra00519g-t1.tif(1)
The basis functions sl,l′;μ(θ,θ′,ϕ) are products of normalized associated Legendre functions Plm:
 
image file: d1ra00519g-t2.tif(2)
where 〈…|…〉 is a Clebsch–Gordan coefficient. The Plm functions are related to spherical harmonics through Ylm(θ,ϕ) = Plm(θ)exp(imϕ). Here, l and l′ are associated with the rotational motion of C2H and para-H2 respectively. In eqn (1) the homonuclear symmetry of H2 forces the index l′ to be even. For collisions at low/moderate temperature, the probability of rotational excitation of H2 is low (the energy spacing between the jH2 = 0 and jH2 = 2 levels in para-H2 being 510 K) so we further restrict H2 to its lowest rotational level. In this case, only the leading term image file: d1ra00519g-t3.tif needs to be retained in the expansion of the interaction potential given in eqn (1). The resulting expansion then can be simplified to
 
image file: d1ra00519g-t4.tif(3)
where the Vav(R,θ) is obtained by an average over angular motion (θ′,ϕ) of the H2 molecule. We approximate the average by an equipoise averaging over three sets of (θ′,ϕ) angles for each calculated set of (R,θ). Those are image file: d1ra00519g-t5.tif, image file: d1ra00519g-t6.tif and image file: d1ra00519g-t7.tif Thus, the C2Hpara-H2(jH2 = 0) is reduced to a two-dimensional Vav PES such as
 
image file: d1ra00519g-t8.tif(4)


image file: d1ra00519g-f1.tif
Fig. 1 Definition of the body-fixed coordinate system for C2H–H2.

Such approximation has been shown to be reasonably accurate for relatively heavy molecule such as SiS,32 HCO+ (ref. 33) and C2.34 In the Cs point group, the ground electronic state of the C2Hpara-H2 van der Waals system is of A′ symmetry. The PES was calculated in the super-molecular approach based on a single and double excitation coupled cluster method with perturbative contributions from triple excitations (CCSD(T)).35,36 For the five atoms, we used the aug-cc-pVTZ basis set of Woon and Dunning.37 This basis set was further augmented by the additional 3s3p2d1f bond basis functions of Williams et al.38 and placed equidistant between the C2H and H2 centers of mass. At all geometries, the Boys and Bernardi39 counterpoise procedure was used to correct for basis set superposition error (BSSE). All calculations were carried out using the MOLPRO 2010 package.40

The radial scattering coordinate R was assigned 36 values ranging from 3.0 to 50.0 bohr, the θ grid ranged from 0° to 180° in steps of 15°. This resulted in a total of 1404 geometries computed for the C2Hpara-H2(jH2 = 0) system.

An analytic representation of the present 2D PES suitable for dynamics calculations was obtained using the procedure described by eqn (12)–(16) of Werner et al.41 for the CN–He system. In order to perform the scattering calculations, this PES was expanded in terms of Legendre polynomials.

Fig. 2 displays the contour plot of the 2D V(R,θ) PES. For this van der Waals system, the global minimum was found to be 250.6 cm−1 at R = 7.20 bohr and θ = 76.0 degree.


image file: d1ra00519g-f2.tif
Fig. 2 Contour plot of the C2H–H2 PES as a function of R and θ.

3 Dynamical calculation

The fitted C2Hpara-H2 2D-PES was used to calculate state-to-state cross sections and rate coefficients. The full close coupling approach first introduced by Arthurs & Dalgarno42 was used for the calculations of the state-to-state cross sections between the 16 first rotational levels. The energies of the rotational levels were computed from the following C2H spectroscopic constants: B0 = 1.3889354 cm−1 and D0 = 3.2345 10−6 cm−1.24 The scattering calculations were carried out with the MOLSCAT code.43 Calculations were performed for energies ranging from 3.0 to 1200.0 cm−1. The integration parameters and the sise of the basis set were chosen to ensure convergence of the cross sections over this range. At the highest considered total energy (1200.0 cm−1), the C2H rotational basis included channels up to j = 28 to ensure convergence of the excitation cross sections for transitions including up to the j = 16 rotational level (see Tables 1 and 2).
Table 2 Convergence of cross-sections (in Å2) for some rotational transitions with respect to size of the basis set (jmax) for total energy E = 1200 cm−1
  jmax = 27 jmax = 28 jmax = 29 jmax = 30 jmax = 32
σ16–15 15.316 15.322 15.322 15.322 15.322
σ16–14 6.666 6.665 6.665 6.665 6.663
σ2–1 2.7937 2.7941 2.7941 2.7941 2.7941


We carefully spanned the energy grid to take into account the presence of resonances. The energy steps are 0.1 cm−1 below 100.0 cm−1, 0.2 cm−1 from 100.0 to 250.0 cm−1, 1.0 cm−1 between 250.0 and 300.0, 5.0 cm−1 between 300.0 and 400.0 cm−1, 10.0 cm−1 between 400.0 and 500.0 cm−1, 25.0 cm−1 between 500.0 and 700.0 cm−1 and 50.0 cm−1 between 700.0 and 1200.0 cm−1.

3.1 Integral cross sections

Fully-quantum close-coupling calculations of integral cross sections were carried out for values of total energies ranging from 3.0 to 1200 cm−1. The variation of integral cross sections for a few selected jj′ rotational transitions with collision energy is shown in Fig. 3. For collisional energies below 300.0 cm−1 many resonances occur. This is a consequence of the attractive potential well with a depth of 252.6 cm−1. At low energy, the molecule can be temporarily trapped in quasi-bound states,44,45 which arise both from tunneling through the centrifugal barrier (shape resonances) and from excitation of C2H to a higher rotational level which is energetically allowed in the well (Feshbach resonances). Some anisotropy appears around the well depth which may induce a different trend of the cross sections at lower energies.
image file: d1ra00519g-f3.tif
Fig. 3 Rotational excitation jj′ cross sections of C2H in collision with para-H2(jH2 = 0) as a function of the relative kinetic energy.

3.2 Rate coefficients

From the calculated cross sections, one can obtain the corresponding state-to-state rate coefficients by Boltzmann averaging:
 
image file: d1ra00519g-t9.tif(5)
where kB is the Boltzmann constant and Ek is the kinetic energy. The total energy range considered in this work allows us to determine rate coefficients for temperatures up to 120 K. The total energy range considered in this work allows us to determine rate coefficients for temperatures up to 120 K. The representative variation with temperature of de-excitation rate coefficients from an initial level j to a final level j′ is shown in Fig. 4. The rate coefficients obviously display the same propensity than the integral cross sections. Fig. 5 present for 50.0 K downward rotational rate coefficients of C2H (j = 12) level as a function of the final j level. This plot confirms the Δj = 2 even propensity. The same behavior was found for the isoelectronic HCN molecule46 and for the neutral C2H47 in collision with para-H2(jH2 = 0); and the C2H anion in collision with He.48

image file: d1ra00519g-f4.tif
Fig. 4 Temperature dependence of state-to-state rate coefficients.

image file: d1ra00519g-f5.tif
Fig. 5 C2Hpara-H2(jH2 = 0) excitation rate coefficients for the jj′ = 12 transitions at 50 K.

The complete set of de-excitation rate coefficients with j, j′ ≤ 16 will be available on-line from the BASECOL website (http://www.obspm.fr/basecol). Excitation rate coefficients can be easily obtained by detailed balance.

4 Comparison with C2H–He collisions

It is generally assumed that excitation rate coefficients associated with He as the collision partner can provide a first estimation for rate coefficients for collisions with para-H2(jH2 = 0) just applying a scaling factor which is equal to the square root of the ratio of the reduced masses of the two systems. For C2H, this scaling factor tends to 1.4. Since rotational rate coefficients have been calculated previously for C2H in collision with He,48 using the same methodology as the one employed in present work for C2Hpara-H2(jH2 = 0), the comparison between the two sets of rate coefficients can be used to assess the validity of such approximation.

We report in Fig. 6 a comparison between the rate coefficients of the present study and those reported by Dumouchel et al.48 for the selected transitions (see Fig. 4 in ref. 48). The red circles and the blue diamond's reveal the rate coefficients of C2H−1para-H2(jH2 = 0) vs. C2H–He at kinetic temperatures T = 10.0 K and T = 100.0 K respectively. As shown above, the ratio between the rate coefficients clearly differs from the value of 1.4. These differences are due to the difference in interaction energies between the two systems. The ratios deviate clearly from 1.4 for all the transitions, by a factor varying from 3 to 11. The difference is more pronounced for the transitions with odd Δj. This is due to the difference in the interaction potential which is more symmetrical (with respect to the θ ↔ π − θ transformation) for the C2H–He system than for C2H−1para-H2(jH2 = 0). Then, the collision rates with corrected with the scale factor 1.4 do not always constitute a good approximation to model the rates with H2. This result is also found in the collisions with the neutral species C2H (see ref. 47).


image file: d1ra00519g-f6.tif
Fig. 6 C2Hpara-H2(jH2 = 0) rate coefficients as a function of the C2H–He rate coefficients for jj′ transitions with j = 12 and j′ = 6, 7, 8, 9, 10, 11 at a temperature of 10 K (red circles) and 100 K (blue diamond); the black solid line corresponds to a ratio of 1.4.

5 Summary and conclusion

Quantum scattering calculations have been used to investigate energy transfer in collisions of C2H with para-H2(jH2 = 0) molecule. The calculations are based on a new, highly correlated 2D PES calculated at the RCCSD(T) level using large AVTZ basis sets and a bend functions. Close coupling calculations were performed for collision energies ranging from 3.0 to 1200.0 cm−1. Rate coefficients for transitions involving the 16 first rotational levels of the C2H anion were determined for temperature ranging from 5 to 120.0 K. The rate coefficients show a strong propensity for even Δj, mainly Δj = 2. The present set of rate coefficients can serve the modelling of C2H emission lines in astrophysical environments. Finally, the resulting inelastic rate coefficients for collisions of C2H with para-H2 will help in constraining astrophysical anion chemistry and also in accurately model regions containing anions such as IRC+10216. We encourage astrophysicists to use these new values in their next detection attempts in the IRC+10216. Future work will deal with calculations of C2H in collision with both para- and ortho-H2 species using 4D PES approach in order to obtain rate coefficients for highly excited rotational states and higher temperatures. Detailed comparison with other systems, in particular carbon chain anions, will be studied in detail owing to its great interest for astrophysical modelling.

Conflicts of interest

There are no conflicts to declare.

Notes and references

  1. C. F. Lind Jabka, M. Polk, I. Zymak and W. D. Geppert, Phys. Chem. Chem. Phys., 2018, 20, 5377–5388 RSC.
  2. M. C. McCarthy, C. A. Gottlieb, H. Gupta and P. Thaddeus, Astrophys. J., 2006, 652, 141–144 CrossRef.
  3. H. Gupta, S. Brunken, F. Tamassia, C. A. Gottlieb, M. C. McCarthy and P. Thaddeus, Astrophys. J., 2007, 655, 57–60 CrossRef.
  4. S. Brunken, H. Gupta, C. A. Gottlieb, M. C. McCarthy and P. Thaddeus, Astrophys. J., 2007, 664, 43–46 CrossRef.
  5. Y. Kasai, E. Kagi and K. Kawaguchi, Astrophys. J., 2007, 661, 61–64 CrossRef.
  6. T. J. Millar, C. Walsh and A. Thomas, Field Chem. Rev., 2017, 117, 17651795 Search PubMed.
  7. M. A. Cordiner, T. J. Millar, C. Walsh, E. Herbst, D. C. Lis, T. A. Bell and E. Roueff, Organic Matter in Space, ed. S. Kwok and S. Sandford, Proceedings IAU Symposium no. 251, 2008 Search PubMed.
  8. M. A. Cordiner and T. J. Millar, Astrophys. J., 2009, 697, 68 CrossRef CAS.
  9. E. Herbst and Y. Osamura, Astrophys. J., 2008, 679, 1670 CrossRef CAS.
  10. C. Walsh, N. Harada, E. Herbst and T. J. Millar, Astrophys. J., 2009, 700, 752 CrossRef CAS.
  11. B. Eichelberger, T. P. Snow, C. Barckholtz and V. M. Bierbaum, Astrophys. J., 2007, 667, 1283 CrossRef CAS.
  12. A. Dalgarno and R. A. Mc Cray, Astrophys. J., 1973, 181, 95 CrossRef CAS.
  13. E. Herbst, Nature, 1981, 289, 656 CrossRef CAS.
  14. K. Kawaguchi, Y. Kasai, S. Ishikawa and N. Kaifu, Publ. Astron. Soc. Jpn., 1995, 47, 853 CAS.
  15. J. Cernicharo, M. Guélin and M. Agúndez, et al., Astron. Astrophys., 2007, 467, L37 CrossRef CAS.
  16. M. Agúndez, J. Cernicharo, M. Guelin, M. Gerin, M. C. McCarthy and P. Thaddeus, Astron. Astrophys., 2008, 478, L19 CrossRef.
  17. N. Sakai, T. Sakai and S. Yamamoto, Astrophys. J., 2008, 673, L71 CrossRef CAS.
  18. A. J. Remijan, J. M. Hollis and F. J. Lovas, et al., Astrophys. J., 2007, 664, L47 CrossRef CAS.
  19. K. Kawaguchi, et al., Publ. Astron. Soc. Jpn., 2007, 59, L47 CrossRef CAS.
  20. S. Brünken, H. Gupta and C. A. Gottlieb, et al., Astrophys. J., 2007, 664, L43 CrossRef.
  21. P. Thadeus, C. A. Gottlieb and H. Gupta, et al., Astrophys. J., 2008, 677, 1132 CrossRef.
  22. C. Walsh, N. Harada, E. Herbst and T. J. Millar, Astrophys. J., 2009, 700, 752 CrossRef CAS.
  23. M. Agúndez, et al., Astron. Astrophys., 2010, 517, L2 CrossRef.
  24. S. Brünken, C. A. Gottlieb and H. Gupta, et al., Astron. Astrophys., 2007, 464, L33 CrossRef.
  25. J. Cernicharo, M. Guélin and M. Agúndez, et al., Astrophys. J., 2008, 688, L83 CrossRef CAS.
  26. M. Agúndez, et al., Astron. Astrophys., 2010, 517, L2 CrossRef.
  27. A. Spielfiedel, N. Feautrier, F. Najar, D. Ben Abdallah, F. Dayou, M. L. Senent and F. Lique, Mon. Not. R. Astron. Soc., 2012, 421, 1891–1896 CrossRef CAS.
  28. X. Huang and T. J. Lee, J. Chem. Phys., 2009, 131, 1043 Search PubMed.
  29. A. Faure, Private Communication, 2011 Search PubMed.
  30. A. Faure, P. Valiron, M. W. L. Wiesenfeld, C. Rist, J. Noga and J. Tennyson, J. Chem. Phys., 2005, 122, 221201 CrossRef PubMed.
  31. T. S. Green, Astrophys. J., 1975, 201, 366 CrossRef.
  32. F. Lique and J. Klos, J. Chem. Phys., 2008, 128, 034306 CrossRef PubMed.
  33. O. Yazidi, D. Ben Abdallah and F. Lique, Mon. Not. R. Astron. Soc., 2014, 441, 664 CrossRef CAS.
  34. F. Najar and Y. Kalugina, RSC Adv., 2020, 10, 8580 RSC.
  35. C. Hampel, K. A. Peterson and H.-J. Werner, Chem. Phys. Lett., 1992, 190, 1 CrossRef CAS.
  36. J. D. Watts, J. Gauss and R. J. Bartlett, J. Chem. Phys., 1993, 98, 8718 CrossRef CAS.
  37. D. E. Woon and T. H. Dunning Jr, J. Chem. Phys., 1994, 100, 2975 CrossRef CAS.
  38. H. L. Williams, E. M. Mas, K. Szalewicz and B. Jeziorski, J. Chem. Phys., 1995, 103, 7374 CrossRef CAS.
  39. S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553 CrossRef CAS.
  40. H.-J. Werner, et al., Molpro, a Package of Ab Initio Programs, 2010, Available at: http://www.molpro.net Search PubMed.
  41. H.-J. Werner, B. Follmeg, M. H. Alexander and D. Lemoine, J. Chem. Phys., 1989, 91, 5425 CrossRef CAS.
  42. A. M. Arthurs and A. Dalgarno, Proc. R. Soc. A, 1960, 256, 540 Search PubMed.
  43. J. M. Hutson and S. Green, MOLSCAT computer code, Version 14, Distributed by Collaborative Computational Project 6, Daresbury Laboratory, Warington, UK, 1995 Search PubMed.
  44. L. N. Smith, D. J. Malik and D. Secrest, J. Chem. Phys., 1979, 71, 4502 CrossRef CAS.
  45. K. M. Christoffel and J. M. Bowman, J. Chem. Phys., 1983, 78, 3952 CrossRef CAS.
  46. D. Ben Abdallah, F. Najar, N. Jaidane, F. Dumouchel and F. Lique, Mon. Not. R. Astron. Soc., 2012, 419, 2441 CrossRef CAS.
  47. F. Najar, D. Ben Abdallah, A. Spielfiedel, F. Dayou, F. Lique and N. Feautrier, Chem. Phys. Lett., 2014, 614, 251–257 CrossRef CAS.
  48. F. Dumouchel, A. Spielfiedel, M. L. Senent and N. Feautrier, Chem. Phys. Lett., 2012, 533, 6–9 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/d1ra00519g

This journal is © The Royal Society of Chemistry 2021